13.07.2015 Views

(GST-3) and ten Delta-class (GST-1)

(GST-3) and ten Delta-class (GST-1)

(GST-3) and ten Delta-class (GST-1)

SHOW MORE
SHOW LESS
  • No tags were found...

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Biochem. J. (2003) 370, 661–669 (Printed in Great Britain) 661Cloning, expression <strong>and</strong> biochemical characterization of one Epsilon-<strong>class</strong>(<strong>GST</strong>-3) <strong>and</strong> <strong>ten</strong> <strong>Delta</strong>-<strong>class</strong> (<strong>GST</strong>-1) glutathione S-transferases fromDrosophila melanogaster, <strong>and</strong> identification of additional nine membersof the Epsilon <strong>class</strong>Rafa SAWICKI*, Sharda P. SINGH*, Ashis K. MONDAL*, Helen BENES † <strong>and</strong> Piotr ZIMNIAK*‡§ 1*Department of Medicine, University of Arkansas for Medical Sciences, Little Rock, AR 72205, U.S.A., †Department of Anatomy <strong>and</strong> Neurobiology, University ofArkansas for Medical Sciences, Little Rock, AR 72205, U.S.A., ‡Department of Biochemistry <strong>and</strong> Molecular Biology, University of Arkansas for Medical Sciences, LittleRock, AR 72205, U.S.A., <strong>and</strong> §Central Arkansas Veterans Healthcare System, Medical Research (151/LR), 4300 West 7th Street, Little Rock, AR 72205, U.S.A.From the fruitfly, Drosophila melanogaster, <strong>ten</strong> members of thecluster of <strong>Delta</strong>-<strong>class</strong> glutathione S-transferases (<strong>GST</strong>s; formerlydenoted as Class I <strong>GST</strong>s) <strong>and</strong> one member of the Epsilon-<strong>class</strong>cluster (formerly <strong>GST</strong>-3) have been cloned, expressed inEscherichia coli, <strong>and</strong> their catalytic properties have been determined.In addition, nine more members of the Epsilon clusterhave been identified through bioinformatic analysis but notfurther characterized. Of the 11 expressed enzymes, seven acceptedthe lipid peroxidation product 4-hydroxynonenal assubstrate, <strong>and</strong> nine were active in glutathione conjugation of 1-chloro-2,4-dinitrobenzene. Since the enzymically active proteinsincluded the gene products of Dm<strong>GST</strong>D3 <strong>and</strong> Dm<strong>GST</strong>D7 whichwere previously deemed to be pseudogenes, we investigated themfurther <strong>and</strong> determined that both genes are transcribed inDrosophila. Thus our present results indicate that Dm<strong>GST</strong>D3<strong>and</strong> Dm<strong>GST</strong>D7 are probably functional genes. The exis<strong>ten</strong>ce <strong>and</strong>multiplicity of insect <strong>GST</strong>s capable of conjugating 4-hydroxynonenal,in some cases with catalytic efficiencies approachingthose of mammalian <strong>GST</strong>s highly specialized for this function,indicates that metabolism of products of lipid peroxidation is ahighly conserved biochemical pathway with probable detoxificationas well as regulatory functions.Key words: detoxification, Diptera, electrophile, glutathioneconjugation, lipid peroxidation.INTRODUCTIONPartially reduced forms of oxygen, including the superoxideradical anion, hydrogen peroxide <strong>and</strong> the hydroxyl radical, arecollectively known as reactive oxygen species (ROS). The formationof ROS is an inevitable consequence of aerobic metabolism.The majority of ROS is generated by mitochondria, althoughother enzyme systems also contribute; transition metal ionscatalyse interconversions of individual ROS, especially the formationof the highly reactive hydroxyl radical. The latter can react,among other targets, with polyunsaturated fatty acids <strong>and</strong> thusinitiate the lipid peroxidation chain reaction [1], which producesmultiple molecules of lipid hydroperoxides per initiating event.By either spontaneous or enzyme-catalysed reactions [2], lipidhydroperoxides give rise to α,β-unsaturated aldehydes such as 4-hydroxynonenal (4-HNE) [3]. These aldehydes are highly reactivebut considerably more selective than the parent ROS. Specificmodification of proteins by 4-HNE <strong>and</strong> similar aldehydesof<strong>ten</strong> affects protein function. This is probably the basis ofthe signalling role of 4-HNE at physiological concentrations[4,5] <strong>and</strong> the toxicity elicited by this compound at highconcentrations [3].The generation of 4-HNE, primarily from arachidonic acid [3],<strong>and</strong> the formation of 4-HNE–protein adducts [6] are wellestablishedphenomena in mammalian tissues. Although thefruitfly, Drosophila melanogaster, lacks arachidonic acid [7,8], 4-HNE can be formed from any ω-6 polyunsaturated fatty acid [9].In addition to arachidonic acid, this also includes linoleic acid,which is present in large amounts in larval <strong>and</strong> adult stages ofDrosophila [8,10]. Although the ability of Drosophila to desaturatefatty acids is limited, dietary polyunsaturated fatty acids arereadily incorporated into phospholipids [8]. Therefore 4-HNE<strong>and</strong>or similar 4-hydroxyalkenals are expected to be generated inDrosophila. We have previously confirmed the presence of 4-HNE–protein adducts in Drosophila by immunoblotting [11],<strong>and</strong> Yan <strong>and</strong> Sohal [12] have demonstrated the formation ofsuch adducts in the housefly.In mammals, the major route of 4-HNE metabolism is viaglutathione conjugation [13], catalysed by a specialized sub<strong>class</strong>of Alpha-<strong>class</strong> glutathione S-transferases (<strong>GST</strong>s) [14,15], althougha reductive pathway may be significant in some tissues.Because of the apparently universal <strong>and</strong> obligatory formation of4-hydroxyalkenals in aerobes, we examined which of the Drosophila<strong>GST</strong>s has the ability to conjugate 4-HNE. In a previousstudy [11], we investigated Dm<strong>GST</strong>S1-1 (see the Materials <strong>and</strong>methods section for an explanation of the nomenclature), whichwas previously assumed to be a structural muscle protein orperhaps a stretch sensor devoid of enzymic properties [16].However, Dm<strong>GST</strong>S1-1 has moderate glutathione-conjugatingactivity for 4-HNE, <strong>and</strong> due to its high level of expression, itaccounts for most of this activity in adult Drosophila [11]. In thecourse of that study, we also found a less abundant <strong>GST</strong> (ormixture of <strong>GST</strong>s), probably belonging to the <strong>Delta</strong> <strong>class</strong>, whichhad a higher specific activity for 4-HNE conjugation thanDm<strong>GST</strong>S1-1 [11]. We have now characterized a number ofDrosophila <strong>GST</strong>s, both of the <strong>Delta</strong> <strong>and</strong> the novel Epsilon <strong>class</strong>,<strong>and</strong> have shown that some, but not all, of them participate in4-HNE metabolism.Abbreviations used: CDNB, 1-chloro-2,4-dinitrobenzene; EST, expressed sequence tag; <strong>GST</strong>, glutathione S-transferase; 4-HNE, 4-hydroxynonenal;ORF, open reading frame; poly(A)+, polyadenylated; ROS, reactive oxygen species; RT, reverse transcriptase.1 To whom correspondence should be addressed (e-mail zimniakpiotruams.edu). 2003 Biochemical Society


662 R. Sawicki <strong>and</strong> othersMATERIALS AND METHODSNomenclature of Drosophila <strong>GST</strong>sIn the present study, we adhere to a recently proposed unifiednomenclature that includes insect as well as mammalian <strong>GST</strong>s[17]. Accordingly, we use the term ‘<strong>Delta</strong> <strong>class</strong>’ for Drosophila<strong>GST</strong>-1 <strong>and</strong> its homologues [18], refer to the protein originallydescribed as <strong>GST</strong>-2 [19] as Dm<strong>GST</strong>S1-1 <strong>and</strong> designate Drosophila<strong>GST</strong>-3 [20] as Dm<strong>GST</strong>E1-1, a prototypical member of theEpsilon <strong>class</strong> of <strong>GST</strong>s (Professor Philip G. Board, personal communication).This nomenclature has been recently adopted byFlybase (www.flybase.org).Cloning of Drosophila <strong>Delta</strong>- <strong>and</strong> Epsilon-<strong>class</strong> <strong>GST</strong>sPCR primers specific for the <strong>ten</strong> <strong>Delta</strong>-<strong>class</strong> genes <strong>and</strong> forDm<strong>GST</strong>E1 were used to amplify the sequences from Drosophilagenomic DNA. The primers are listed in Table 1. All upstreamprimers were designed to span the translation initiation codon<strong>and</strong> to introduce an NdeI restriction site to facilitate subsequentsubcloning. For the same reason, the downstream primersintroduced a unique restriction site, as specified in Table 1. Onehundred male adult flies were homogenized, <strong>and</strong> total genomicDNA was isolated using the DNeasy Tissue Kit (Qiagen). EachPCR reaction contained, in a total volume of 25 µl, 50 ng ofDrosophila genomic DNA as the template <strong>and</strong> 25 ng each of therespective primers. Reactions were initiated by the addition ofTaq polymerase using the hot start technique <strong>and</strong> were continuedfor 32 cycles of denaturation for 30 s at 95 C, annealing for 30 sat 48 C for Dm<strong>GST</strong>D6, 50C for Dm<strong>GST</strong>D10, ramped from52.2 to 49 C over 32 cycles (touch-down PCR) for Dm<strong>GST</strong>D7<strong>and</strong> from 56.4 to 50 C for Dm<strong>GST</strong>D1–5, Dm<strong>GST</strong>D8, Dm<strong>GST</strong>D9<strong>and</strong> Dm<strong>GST</strong>E1, followed by ex<strong>ten</strong>sion for 1 min at 72 Ccycle.The specificity of the amplification was verified by sequencing ofall PCR products.Bacterial expression <strong>and</strong> purification of Drosophila <strong>GST</strong>sThe PCR-amplified fragments encoding the Drosophila <strong>GST</strong>swere digested with NdeI <strong>and</strong> a restriction enzyme appropriate forthe antisense primer (see Table 1) <strong>and</strong> subcloned between thecorresponding sites of pET-30a() (Novagen). Escherichia coliBL21(DE3)pLysS cells were used for expression of all proteinsexcept for Dm<strong>GST</strong>D3-3, Dm<strong>GST</strong>D7-7 <strong>and</strong> Dm<strong>GST</strong>E1-1, whichwere expressed in the BL21 Star(DE3)pLysS strain, which carriesa truncated RNAse E, leading to increased mRNA stability.Expression was induced with 1 mM isopropyl β-D-thiogalactosideat absorbance A 0.4–0.6 <strong>and</strong> it was continued for 8 h at37 C. The bacteria were collected by centrifugation <strong>and</strong> thepellets were frozen at 20 C until use. Bacterial cells were lysed<strong>and</strong> the <strong>GST</strong>s, except for Dm<strong>GST</strong>D3-3 <strong>and</strong> Dm<strong>GST</strong>E1-1, werepurified by glutathione affinity chromatography using a 1 mlprepacked glutathione Sepharose 4 Fast Flow column(Amersham) according to the manufacturer’s instructions.Dm<strong>GST</strong>D3-3 <strong>and</strong> Dm<strong>GST</strong>E1-1 did not bind to the glutathionecolumn. The bacterial lysate (from 100 ml culture) containingDm<strong>GST</strong>D3-3 was fractionated by ammonium sulphate precipitation.The 25–50% fraction was dialysed against 10 mMpotassium phosphate (pH 7.4) <strong>and</strong> further fractionated on aSephadex G-200 column (1 cm50 cm). Fractions containingthe <strong>GST</strong> were pooled <strong>and</strong> loaded on a hydroxyapatite column(1 cm8.5 cm). The column was eluted with a gradient of10–400 mM potassium phosphate (pH 6.5). Fractions containingDm<strong>GST</strong>D3-3 eluted at 300 mM KCl <strong>and</strong> were identified bySDSPAGE. Dm<strong>GST</strong>E1-1 was purified by ion-exchange chromatography.Frozen bacterial cells from 500 ml of culture werelysed by sonication in 20 mM TrisHCl (pH 7.5), 50 mM KCl,centrifuged for 30 min at 30000 g <strong>and</strong> the supernatant waspassed over a DEAE column (Macro-Prep DEAE Support; Bio-Rad Laboratories), which did not retain Dm<strong>GST</strong>E1-1. The flowthroughwas then applied to a Macro-Prep High Q Supportcolumn (0.5 cm3 cm; Bio-Rad Laboratories), which waswashed with 20 mM TrisHCl (pH 7.5), 50 mM KCl <strong>and</strong> elutedwith a step gradient of KCl in 20 mM TrisHCl (pH 7.5). The4-HNE-conjugating activity eluted at 0.1 M KCl. The fractionwas dialysed against 50 mM TrisHCl (pH 7.5) <strong>and</strong> 1.3 mM2-mercaptoethanol.Determination of enzyme activity<strong>GST</strong> activities were measured spectrophotometrically in a microtitreplate reader (SpectraMax Plus; Molecular Devices, Sunnyvale,CA, U.S.A.). Conjugation of 1-chloro-2,4-dinitrobenzene(CDNB) was measured at 25 C. Enzyme activity with 4-HNEwas calculated from the rate of consumption of this substrate at30 C [13]. To determine kinetic constants of 4-HNE conjugation,the <strong>GST</strong> activities of the enzymes were measured by varying theconcentrations of 4-HNE at fixed concentrations of reducedglutathione. The results were analysed by least-squares nonlinearfitting of a Michaelis–Men<strong>ten</strong> hyperbola or the HillTable 1Primers used to amplify D. melanogaster <strong>GST</strong> genesForward primers were designed to introduce an NdeI restriction site (underline) which includes the initiation codon ATG. Reverse primers spanned all or part of the stop codon (double underline)<strong>and</strong> were designed to introduce a unique restriction site (lower-case boldface; EcoRI for Dm<strong>GST</strong>D1, Dm<strong>GST</strong>D3, Dm<strong>GST</strong>D5, Dm<strong>GST</strong>D7–9 <strong>and</strong> Dm<strong>GST</strong>E1; BamHI for Dm<strong>GST</strong>D2 <strong>and</strong> Dm<strong>GST</strong>D10;KpnI for Dm<strong>GST</strong>D4; SacI for Dm<strong>GST</strong>D6).Clone Forward primer Reverse primer<strong>GST</strong>D1 5-CTTCTACAGTCATATGGTTGACTTCTACTAC 5-AAAACGTgaattcCAGGCTTATTC<strong>GST</strong>D2 5-GCAATATCCCCCATATGGACTTTTACTAC 5-ATAGTTGggatccCGATCACTTAGC<strong>GST</strong>D3 5-CGCTCCGTTCATATGGTGGGCAAG 5-TCgaattcAATGAATTATTTAGCAGCATTCTG<strong>GST</strong>D4 5-CGCCCCAGCCATATGGATTTCTACT 5-GTTTggtaccACTGTAGTTACTTTAATGAG<strong>GST</strong>D5 5-ACACTCAAATAACCCATATGGATTTCTATT 5-TTgaattcTTATTGCTTCGCCGCC<strong>GST</strong>D6 5-GCCCTTCTGCATATGGATCTCTATA 5-CGgagctcTCTATAATATTTATAACTTTTC<strong>GST</strong>D7 5-GTGAGTTCTACATATGACAAACATATTC 5-GTgaattcAAAACTATTCTACGGTG<strong>GST</strong>D8 5-GGACCTTTCCATATGGACTTTTACTACC 5-CATTTACTGCAGgaattcTCATCAAATC<strong>GST</strong>D9 5-CAGCGAGTACATATGTTGGACTTCTAC 5-GATgaattcTTACAGACCCGTCATTG<strong>GST</strong>D10 5-GAGATAGCGTCATATGGATTTATACTATAG 5-TCGggatccTATATTCAGTTATCGAAG<strong>GST</strong>E1 5-ACAACACATCATATGTCGAGCTCTGG 5-TTgaattcACTTGTCGCCCAGC 2003 Biochemical Society


Cloning <strong>and</strong> expression of Drosophila glutathione S-transferases663equation. Glutathione peroxidase activity with cumene hydroperoxideas substrate was determined as described in [21].Reverse transcriptase (RT)–PCRThe presence of Dm<strong>GST</strong>D3 transcripts was determined in totalRNA isolated from w<strong>and</strong>ering third-instar Drosophila larvaeusing the RNeasy Mini Kit (Qiagen). The RNA was treated withRNase-free DNase I (Roche Applied Science, Indianapolis, IN,U.S.A.; 1 unit of DNase Iµg of RNA in 25 mM TrisHCl,5 mM MgCl , 0.1 mM EDTA, at 37 C for 30 min, followed byheat inactivation of DNase at 75 C for 15 min). RNA (200 ng)was used for each RT–PCR reaction (Titan One Tube RT–PCRSystem; Roche Applied Science). The sense primer was 5-CGCTCCGTTCTGATGGTGGGCAAG, <strong>and</strong> the antisenseprimer was 5-TCGTTATAAATGAATTATTTAGCAGCAT-TCTG. The reverse transcription step was performed at 50 Cfor 30 min, followed by inactivation of the RT at 94 C for 2 min.The conditions for the subsequent PCR step were as follows: 10cycles of 94 C, 10 s; 58 C, 30 s; 68 C, 45 s <strong>and</strong> then 25 cyclesof 94 C, 10 s; 58 C, 30 s; 68 C, initially 45 s with cycleelongation by 5 s for each cycle; 1 cycle at 94 C, 10 s; 58 C,30 s; 68 C for 7 min. In control reactions, the mixture of RT <strong>and</strong>thermostable DNA polymerase supplied in the Titan One TubeRT–PCR System was substituted with Taq polymerase (Biolase;Bioline USA, Canton, MA, U.S.A.). For detection of theDm<strong>GST</strong>D7 transcript, polyadenylated [poly(A)+] RNA was isolatedfrom 25 µg of total larval RNA using the Oligotex DirectmRNA Mini Kit (Qiagen), <strong>and</strong> one-<strong>ten</strong>th of each of the recoveredpoly(A)+ RNA was used for every RT–PCR or control reactionunder conditions described above for Dm<strong>GST</strong>D3, with the senseprimer 5-GTGAGTTCTAACAATGACAAACATATTC <strong>and</strong>the antisense primer 5-GTCAAACGAAAACCATTCTACG-GTG.RESULTSIdentification of gene clusters encoding <strong>Delta</strong>- <strong>and</strong> Epsilon-<strong>class</strong><strong>GST</strong>s in Drosophila genomeThe genomic sequence of D. melanogaster was examined, usingthe BLAST software, for the presence of open reading frames(ORFs) homologous with previously described <strong>Delta</strong> [18] <strong>and</strong>Epsilon [20] <strong>GST</strong>s. The <strong>Delta</strong> cluster at chromosomal position87B11–87B13 spanned approx. 18 kb <strong>and</strong> contained the eightgenes previously designated as D1, D21, D22, D23, D24, D25,D26 <strong>and</strong> D27 [18,22,23]. In keeping with a recent nomenclatureproposal [17], we refer to these genes as Dm<strong>GST</strong>D1 to Dm<strong>GST</strong>D8respectively. In addition, the <strong>Delta</strong> cluster contained two recentlyreported [17] but not further characterized genes Dm<strong>GST</strong>D9<strong>and</strong> Dm<strong>GST</strong>D10. Three genes (Dm<strong>GST</strong>D1, Dm<strong>GST</strong>D9 <strong>and</strong>Dm<strong>GST</strong>D10) were located on the same DNA str<strong>and</strong>, whereas theremaining <strong>Delta</strong>-<strong>class</strong> genes were on the opposite str<strong>and</strong>. Theorientation <strong>and</strong> relative positions of the <strong>Delta</strong> ORFs are depictedin Figure 1(A).A novel Theta <strong>class</strong>-related <strong>GST</strong> gene Gst-3 has been describedrecently [20]. According to a suggestion from Professor PhilipG. Board (personal communication), we refer to this gene asDm<strong>GST</strong>E1, a member of the Epsilon <strong>class</strong> of <strong>GST</strong>s. By BLASTsearches of the Drosophila genomic sequence, we identified nineadditional genes (Dm<strong>GST</strong>E2–Dm<strong>GST</strong>E10) with high homologyto Dm<strong>GST</strong>E1. The Epsilon genes form a tight cluster spanningapprox. 13 kb at chromosomal position 55C9. Within that cluster(Figure 1B), one gene (Dm<strong>GST</strong>E10) has an antiparallel orientationrelative to the remaining genes.Figure 1 Clusters of <strong>Delta</strong>- <strong>and</strong> Epsilon-<strong>class</strong> glutathione transferasegenes in D. melanogasterThe map is drawn to scale <strong>and</strong> is based on the Drosophila genomic sequence. The arrowsdenote the positions <strong>and</strong> 5 3 direction of the coding sequences of the genes; none of the<strong>Delta</strong> <strong>and</strong> Epsilon genes contains introns within the ORF. The chromosomal positions of theclusters are indicated.Table 2 Context of the initiation codon of D. melanogaster <strong>Delta</strong> <strong>and</strong>Epsilon <strong>GST</strong> genesGene 4 3 2 1 Start 4Dm<strong>GST</strong>D1 T A A A ATG GDm<strong>GST</strong>D2 C A A C ATG GDm<strong>GST</strong>D3 T C T G ATG GDm<strong>GST</strong>D4 C A A C ATG GDm<strong>GST</strong>D5 C G A A ATG GDm<strong>GST</strong>D6 G A C G ATG GDm<strong>GST</strong>D7 A A C A ATG ADm<strong>GST</strong>D8 C A T C ATG GDm<strong>GST</strong>D9 A A T C ATG TDm<strong>GST</strong>D10 T A A G ATG GDm<strong>GST</strong>E1 T A T C ATG TDm<strong>GST</strong>E2 C A T C ATG TDm<strong>GST</strong>E3 A G A C ATG GDm<strong>GST</strong>E4 G A G A ATG GDm<strong>GST</strong>E5 A A A C ATG GDm<strong>GST</strong>E6 C A A G ATG GDm<strong>GST</strong>E7 C A A G ATG CDm<strong>GST</strong>E8 C G G C ATG TDm<strong>GST</strong>E9 A G C G ATG GDm<strong>GST</strong>E10 C A C A ATG GDrosophilaConsensus [31] 3% G 13% G 9% G 18% G ATG 26% G32% A 82% A 56% A 38% A ATG 37% A8% T 1% T 10% T 8% T ATG 22% T57% C 3% C 25% C 36% C ATG 15% CSequence similarities within the <strong>Delta</strong> <strong>and</strong> Epsilon clusters of<strong>GST</strong>sIt has been previously reported that the coding sequences of the<strong>Delta</strong>-<strong>class</strong> genes [18,24,25] as well as of the Dm<strong>GST</strong>E1 gene [20]contain no introns. We have now ex<strong>ten</strong>ded this observation toDm<strong>GST</strong>D9, Dm<strong>GST</strong>D10 <strong>and</strong> the nine novel Dm<strong>GST</strong>E genes. Allof the <strong>Delta</strong> <strong>and</strong> Epsilon genes can be conceptually translated.Their initiation codons are in a context appropriate for Drosophila[26] (Table 2) <strong>and</strong> they are followed by an ORF of a sizeexpected for glutathione transferases (see the Discussion sectionfor additional details). The alignment of the resulting proteinsequences is shown in Figure 2. The <strong>Delta</strong>-<strong>class</strong> sequences formeda tight cluster, with the exception of Dm<strong>GST</strong>D7-7, which carrieda 47-amino-acid N-terminal ex<strong>ten</strong>sion <strong>and</strong> was truncated at theC-terminus by approx. 50 amino acids, <strong>and</strong> of Dm<strong>GST</strong>D3-3,which was shorter by approx. 15 residues at the N-terminus,as compared with the remaining <strong>Delta</strong>-<strong>class</strong> proteins. The 2003 Biochemical Society


664 R. Sawicki <strong>and</strong> othersFigure 2Alignment of the predicted protein sequences of D. melanogaster <strong>Delta</strong>- <strong>and</strong> Epsilon-<strong>class</strong> glutathione transferasesResidues identical in at least 18 of the 20 sequences are shaded in black, <strong>and</strong> residues similar (as defined by the BLOSUM62 matrix) in at least 18 of the 20 sequences are shown as blackletters shaded in light grey. Residues that are, by the above criteria, neither identical nor similar in the entire set of 20 sequences but are similar (or identical) in at least 9 out of 10 sequencesin one of the two <strong>GST</strong> clusters are shown as white letters shaded in dark grey. Thus amino acid positions denoted as white letters shaded in dark grey are conserved within the <strong>Delta</strong> or the Epsiloncluster, but not between the two clusters. 2003 Biochemical Society


Cloning <strong>and</strong> expression of Drosophila glutathione S-transferases665Figure 3 Percentage of identity/similarity respectively for pairwise alignmentsof Drosophila <strong>Delta</strong>-<strong>class</strong> protein sequencesThe PAM250 matrix was used to calculate similarity scores.similarities <strong>and</strong> identities of all pairs of the <strong>Delta</strong>-<strong>class</strong> sequencesare presented in Figure 3. Similarly, all Epsilon-<strong>class</strong> proteinscould be brought into close alignment, with Dm<strong>GST</strong>E10-10 beingthe most divergent sequence because of a C-terminal ex<strong>ten</strong>sionof approx. 15 amino acids (Figure 2). The similarity between the<strong>Delta</strong> <strong>and</strong> Epsilon clusters was less pronounced. For example,the similarity <strong>and</strong> identity between the Dm<strong>GST</strong>D1-1 <strong>and</strong>Dm<strong>GST</strong>E1-1 proteins was only 36 <strong>and</strong> 49% respectively. Therelationship between the two <strong>GST</strong> <strong>class</strong>es can be visualized as ahypothetical phylogenetic tree, which shows that the <strong>Delta</strong> <strong>and</strong>Epsilon nucleotide sequences form distinct clusters, with bothclusters only distantly related to the remaining characterizedDrosophila <strong>GST</strong> gene, Dm<strong>GST</strong>S1 (Figure 4). The sequencehomology within each of the two clusters, together with thephysical proximity of all <strong>Delta</strong> genes on chromosome 3 <strong>and</strong> allEpsilon genes on chromosome 2 (Figure 1), suggests that eachcluster was probably formed by repeated duplication events of a<strong>Delta</strong> <strong>and</strong> an Epsilon ancestral gene respectively. Interestingly, inthe Epsilon cluster, similarities in sequence <strong>and</strong> thus probableevolutionary relationships of individual genes are reflected intheir physical proximity. For example, the Dm<strong>GST</strong>E5 <strong>and</strong>Dm<strong>GST</strong>E6 genes, which have probably diverged most recentlyon an evolutionary time scale (Figure 4), are located next to eachother on the genomic DNA (Figure 1). This generalization holdsfor all members of the Epsilon group of genes, but not for the<strong>Delta</strong> cluster. This indicates that, in the Epsilon cluster, individualgenes (rather than blocks of genes) underwent duplication, withno subsequent rearrangement.Cloning <strong>and</strong> bacterial expression of <strong>Delta</strong>- <strong>and</strong> Epsilon-<strong>class</strong> <strong>GST</strong>sThe <strong>ten</strong> <strong>Delta</strong>-<strong>class</strong> genes (Dm<strong>GST</strong>D1–Dm<strong>GST</strong>D10) as well asDm<strong>GST</strong>E1 were amplified by PCR using genomic DNA as thetemplate, as described in Materials <strong>and</strong> methods section. Theirintron-free coding sequences were used directly for expression ofthe proteins in E. coli. All proteins could be expressed, albeit withdifferent yields. Although the expression level of Dm<strong>GST</strong>D5-5<strong>and</strong> Dm<strong>GST</strong>D7-7 was low (results not shown), it was sufficientto determine the enzymic activity of the proteins in bacteriallysates. Enzymes selected because of their ability to conjugate4-HNE (see below) were further purified (Figure 5).Figure 4Phylogenetic tree of D. melanogaster <strong>GST</strong>sThe tree representing similarities among the full-length protein-encoding nucleotide sequenceswas calculated using the KITSCH program [49], as implemented in the BioEdit program suite(www.mbio.ncsu.edu/BioEdit). The program assumes a constant evolutionary rate; the barindicates the time in which 0.1 nucleotide substitution occurs at any given site.Catalytic activity of Drosophila <strong>GST</strong>sExcept for Dm<strong>GST</strong>D3-3, all <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s conferred CDNBconjugatingactivity on lysates of bacterial cells in which theywere expressed. In contrast, Dm<strong>GST</strong>D3-3 <strong>and</strong> Dm<strong>GST</strong>E1-1 hadno activity with CDNB but were able to conjugate 4-HNE incrude bacterial lysates. This pattern of activities resembles that ofthe quantitatively predominant Drosophila <strong>GST</strong>, Dm<strong>GST</strong>S1-1([11] <strong>and</strong> Table 3). Six of the 10 <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s had 4-HNEconjugatingactivity. Thus each of the 11 enzymes under studywas active with either 4-HNE or CDNB. Enzymes able tocatalyse the conjugation of 4-HNE were of special interest in thecontext of the present study <strong>and</strong> were selected for purification<strong>and</strong> further characterization. The kinetic parameters of the latter<strong>GST</strong>s are shown in Table 3.In this group, Dm<strong>GST</strong>D1-1 had thehighest catalytic efficiency for 4-HNE conjugation, only 4-foldlower than the catalytic efficiency of the highly specialized murineenzyme m<strong>GST</strong>A4-4 [21]. Dm<strong>GST</strong>D7-7 had an appreciablecatalytic efficiency for 4-HNE which exceeded that of Dm<strong>GST</strong>S1-1. The remaining four <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s had lower but clearlymeasurable activities for 4-HNE. As observed by us previously[11], Dm<strong>GST</strong>S1-1 contributes most of 4-HNE-conjugating activityof adult Drosophila because of the high abundance of theprotein. Our present results indicate that Dm<strong>GST</strong>D1-1 <strong>and</strong>several additional <strong>GST</strong>s have higher or equal catalytic-centreactivities <strong>and</strong>or catalytic efficiencies compared with Dm<strong>GST</strong>S1- 2003 Biochemical Society


Cloning <strong>and</strong> expression of Drosophila glutathione S-transferases667Figure 6 Identification of Dm<strong>GST</strong>D3 <strong>and</strong> Dm<strong>GST</strong>D7 transcripts in Drosophilalarvae by RT–PCRTotal RNA was used as the starting material for amplification of Dm<strong>GST</strong>D3, <strong>and</strong> poly(A)+ RNAwas used to amplify Dm<strong>GST</strong>D7. The identity of the amplification products (denoted by arrows)was confirmed by partial sequencing; the additional b<strong>and</strong>s were probably the result ofmispriming <strong>and</strong> were not further analysed. Control reactions lacking RT (lanes labelled RT)did not yield amplification products.(results not shown). Therefore the poly(A)+ RNA fraction waspurified from total RNA for amplification of the latter sequence.The results of RT–PCR demonstrate that the Dm<strong>GST</strong>D3 <strong>and</strong>Dm<strong>GST</strong>D7 genes are expressed in Drosophila larvae, althoughthe level of the Dm<strong>GST</strong>D7 mRNA may be lower than that ofDm<strong>GST</strong>D3.DISCUSSIONThe cluster of D. melanogaster <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s has beenoriginally described, <strong>and</strong> several of the enzymes have beenbiochemically characterized in a series of elegant papers byC.-P. D. Tu et al. [18,22–24,28,29]. Two additional membersof the cluster have been identified recently by bioinformaticsapproaches [17] but have not been biochemically characterized.Similarly, the Dm<strong>GST</strong>E1 (Gst-3) gene has been reported [20], butthe functional properties of the protein remained unknown. Wehave now cloned, expressed in E. coli <strong>and</strong> characterized all <strong>ten</strong><strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s <strong>and</strong> Dm<strong>GST</strong>E1-1. Furthermore, we identified,but have not further characterized, nine additional members ofthe Epsilon cluster. Thus both the <strong>Delta</strong>- <strong>and</strong> the Epsilon-<strong>class</strong><strong>GST</strong>s form families of at least <strong>ten</strong> members each, probably as theresult of repeated gene duplications.Dm<strong>GST</strong>E1-1 <strong>and</strong> all <strong>ten</strong> <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s could be expressedin E. coli as enzymically active proteins. This result appears tocontradict the earlier conclusion that Dm<strong>GST</strong>D3 <strong>and</strong> Dm<strong>GST</strong>D7are likely to be pseudogenes [18]. Dm<strong>GST</strong>D3 was presumed to bea pseudogene on the basis of a sub-optimal Kozak context of thetranslation-initiation codon, which may prevent its synthesis,<strong>and</strong> a 15-amino-acid N-terminal truncation, which may renderthe protein inactive even if synthesized [18]. Our demonstrationof bacterial expression of Dm<strong>GST</strong>D3-3 does not address thequestion of whether m<strong>GST</strong>D3 mRNA can be efficiently translatedin eukaryotic, <strong>and</strong> specifically Drosophila, cells. However, thedatabase of all known <strong>and</strong> predicted Drosophila transcripts ([30];file available as ‘nagadfly.dros.RELEASE2’ on www.fruitfly.org) contains, in addition to Dm<strong>GST</strong>D3, at least twoother sequences that have an identical context of the initiatorATG (TCTG ATG G) <strong>and</strong> encode identifiable proteins: a Zndependentexopeptidase (Flybase symbol CG10073) <strong>and</strong> a celladhesion protein (CG5550). This is consis<strong>ten</strong>t with the fact thatthe Kozak consensus is not an absolute requirement for translation,but rather one of several interdependent factors that codetermineinitiation efficiency. Moreover, Drosophila appears totolerate more deviation from the ‘ideal’ context of the initiatorcodon than mammalian cells [26,31]. In fact, suboptimal determinantsof protein synthesis initiation may render the processmore amenable to regulation [32], an aspect that may be relevantfor some of the <strong>GST</strong>s (see below). An expressed sequence tag(EST) clone (LP11313, GenBank accession nos. AI297103 <strong>and</strong>AY118350) has been isolated that fully matches the genomicDm<strong>GST</strong>D3 sequence, except for the presence of a poly(A)+tail at the 3-end of the EST clone. Finally, we identified aDm<strong>GST</strong>D3 transcript by RT–PCR. This demonstrates thatthe Dm<strong>GST</strong>D3 gene is expressed, at least up to the stage oftranscription <strong>and</strong> transcript processing. This argues against thepossibility that Dm<strong>GST</strong>D3 is a pseudogene.As noted previously [18], Dm<strong>GST</strong>D3-3 lacks the N-terminalsequence (cf. Figure 2) that includes the highly conserved tyrosineresidue involved in the catalytic cycle of <strong>GST</strong>s. Strikingly,Dm<strong>GST</strong>D3-3 is active towards 4-HNE, with kinetic parameterssimilar to those of other <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s (Table 3). Thisindicates that other residues may take the place of the active-sitetyrosine.Dm<strong>GST</strong>D7 was deemed to be a pseudogene because of a C-terminal truncation which may affect the electrophile-bindingpocket of the active site <strong>and</strong>or protein stability [18]. We havedemonstrated by RT–PCR the presence of Dm<strong>GST</strong>D7 mRNA inDrosophila larvae. Furthermore, our results show that theDm<strong>GST</strong>D7-7 protein is active towards both 4-HNE <strong>and</strong> CDNB(Table 3). In fact, its catalytic efficiency for 4-HNE was secondhighest among the enzymes tested in the present study. Theexamples of Dm<strong>GST</strong>D3-3 <strong>and</strong> Dm<strong>GST</strong>D7-7 show that <strong>GST</strong>scan tolerate significant deletions, probably through compensatorychanges in the folding of the polypeptide.It is noteworthy that six out of the <strong>ten</strong> <strong>Delta</strong>-<strong>class</strong> <strong>GST</strong>s, aswell as Dm<strong>GST</strong>E1-1, had significant activity for 4-HNE conjugation.Together with the previously characterized Dm<strong>GST</strong>S1-1 [11], there are at least seven distinct <strong>GST</strong>s in Drosophila withthis activity. The catalytic efficiency of one of the <strong>Delta</strong>-<strong>class</strong>enzymes, Dm<strong>GST</strong>D1-1, was lower by only a factor of 4 whencompared with that of mammalian Alpha-<strong>class</strong> <strong>GST</strong>s highlyspecialized for 4-HNE conjugation, exemplified by m<strong>GST</strong>A4-4[21]. At least three of the enzymes, namely Dm<strong>GST</strong>D3-3,Dm<strong>GST</strong>E1-1 (Table 3) <strong>and</strong> Dm<strong>GST</strong>S1-1 [11], are active towards4-HNE but not towards the near-universal <strong>GST</strong> model substrateCDNB. Finally, 4-HNE-conjugating activity probably arose onmultiple independent occasions during evolution, since it isassociated with some, but not all, of the Theta-related enzymesin Drosophila [Dm<strong>GST</strong>S1-1 [11] <strong>and</strong> certain <strong>Delta</strong>- <strong>and</strong> Epsilon<strong>class</strong><strong>GST</strong>s (this work)], with some, but again not all, Alpha-<strong>class</strong>mammalian <strong>GST</strong>s [15,33] <strong>and</strong> with at least one Pi-<strong>class</strong> <strong>GST</strong> inCaenorhabditis elegans [34]. Collectively, these findings suggestthat conjugation of 4-HNE evolved, <strong>and</strong> is maintained, byselective pressure, rather than being an adventitious activity ofsome <strong>GST</strong>s.In addition to toxicity at supraphysiological levels, 4-HNE hassignalling functions affecting, among others, cell proliferation,differentiation <strong>and</strong> apoptosis (see e.g. [35–39]). It is tempting tospeculate that the multiplicity of <strong>GST</strong>s involved in 4-HNE 2003 Biochemical Society


668 R. Sawicki <strong>and</strong> othersmetabolism is necessary to modulate <strong>and</strong> terminate thesefunctions of 4-HNE in a variety of physiological situations,perhaps including embryogenesis <strong>and</strong> insect metamorphosis. Thelack of Dm<strong>GST</strong>D3, Dm<strong>GST</strong>D5, Dm<strong>GST</strong>D7 <strong>and</strong> Dm<strong>GST</strong>D8transcripts detectable by Northern blotting in adult Drosophila[18] may indicate that their expression is spatially <strong>and</strong> temporallyrestricted, perhaps to pre-adult stages. For example, the ESTrepresenting Dm<strong>GST</strong>D3 (LP11313; Berkeley DrosophilaGenome Project, direct submission to GenBank) has beenisolated from a larvalearly pupal stage of Drosophila, <strong>and</strong> wehave demonstrated the presence of Dm<strong>GST</strong>D3 <strong>and</strong> Dm<strong>GST</strong>D7transcripts in third instar larvae. In addition to transcriptionalregulation, the suboptimal Kozak context of the translationalstart site of some of the above <strong>GST</strong>s is consis<strong>ten</strong>t with translationalregulation. Low <strong>and</strong> tightly controlled expression ischaracteristic of regulatory proteins, but would be unusual for anenzyme whose primary function is general detoxification.Previously, insect <strong>GST</strong>s have been studied primarily becauseof their possible involvement in the metabolic inactivation ofinsecticides [40]. Our work indicates that several Drosophila<strong>GST</strong>s participate in defence mechanisms against oxidative stress,<strong>and</strong> in the metabolism of endogenously formed lipid peroxidationproducts. In this, they functionally resemble mammalian Alpha<strong>class</strong><strong>GST</strong>s [41,42]. A possible role of lipid peroxidation products,<strong>and</strong> thus <strong>GST</strong>s that metabolize them, in the modulation ofembryogenesis <strong>and</strong> development has been already discussed.Another process in which oxidative stress plays a key role isorganismal aging [43,44]. In this context, it is intriguing thatlifespan in Drosophila correlates with the expression level ofDm<strong>GST</strong>S1-1 [45] <strong>and</strong> Dm<strong>GST</strong>D1-1 [46], two enzymes identifiedby us as capable of metabolizing 4-HNE.The above examples demonstrate that the physiological rolesof <strong>GST</strong>s transcend simple detoxification <strong>and</strong> affect a number offundamental biological processes. Drosophila is well suited to thestudy of these processes because of its known genetics <strong>and</strong>availability of methods for molecular interventions. Since basicbiological processes are of<strong>ten</strong> conserved between invertebratessuch as Drosophila <strong>and</strong> mammals, including humans, we expectthat the elucidation of the physiological functions of Drosophila<strong>GST</strong>s will have broad biological implications.We thank Mr Venu Varma for computational help with the identification, in thedatabase of genes po<strong>ten</strong>tially expressed in Drosophila, of transcripts with asuboptimal Kozak context which corresponds to that of Dm<strong>GST</strong>D3. This investigationwas supported in part by the National Institute of Aging (USPHS grant no.AG018845) <strong>and</strong> the National Institute of Environmental Health Sciences (grantno. ES07804).REFERENCES1 Halliwell, B. <strong>and</strong> Gutteridge, J. M. C. (1999) Free Radicals in Biology <strong>and</strong> Medicine,Oxford University Press, Oxford2 Schneider, C., Tallman, K. A., Porter, N. A. <strong>and</strong> Brash, A. R. (2001) Two distinctpathways of formation of 4-hydroxynonenal: mechanisms of nonenzymatictransformation of the 9- <strong>and</strong> 13-hydroperoxides of linoleic acid to 4-hydroxyalkenals.J. Biol. Chem. 276, 20831–208383 Esterbauer, H. (1993) Cytotoxicity <strong>and</strong> genotoxicity of lipid-oxidation products. Am. J.Clin. Nutr. 57, 779S–786S4 Dianzani, M. U. (1998) 4-Hydroxynonenal <strong>and</strong> cell signalling. Free Radic. Res. 28,553–5605 Uchida, K., Shiraishi, M., Naito, Y., Torii, Y., Nakamura, Y. <strong>and</strong> Osawa, T. (1999)Activation of stress signaling pathways by the end product of lipid peroxidation.4-Hydroxy-2-nonenal is a po<strong>ten</strong>tial inducer of intracellular peroxide production.J. Biol. Chem. 274, 2234–22426 Uchida, K. <strong>and</strong> Stadtman, E. R. (1992) Modification of histidine residues in proteinsby reaction with 4-hydroxynonenal. Proc. Natl. Acad. Sci. U.S.A. 89, 4544–45487 Stark, W. S., Lin, T. N., Brackhahn, D., Christianson, J. S. <strong>and</strong> Sun, G. Y. (1993)Fatty acids in the lipids of Drosophila heads: effects of visual mutants, caro<strong>ten</strong>oiddeprivation <strong>and</strong> dietary fatty acids. Lipids 28, 345–3508 Draper, H. H., Philbrick, D. P., Agarval, S., Meidiger, R. <strong>and</strong> Phillips, J. P. (2000)Avid uptake of linoleic acid <strong>and</strong> vitamin E by Drosophila melanogaster. Nutr. Res. 20,113–1209 Pryor, W. A. <strong>and</strong> Porter, N. A. (1990) Suggested mechanisms for the production of4-hydroxy-2-nonenal from the autoxidation of polyunsaturated fatty acids. Free Radic.Biol. Med. 8, 541–54310 Jones, H. E., Harwood, J. L., Bowen, I. D. <strong>and</strong> Griffiths, G. (1992) Lipid compositionof subcellular membranes from larvae <strong>and</strong> prepupae of Drosophila melanogaster.Lipids 27, 984–98711 Singh, S. P., Coronella, J. A., Benes , H., Cochrane, B. J. <strong>and</strong> Zimniak, P. (2001)Catalytic function of Drosophila melanogaster glutathione S-transferase Dm<strong>GST</strong>S1-1(<strong>GST</strong>-2) in conjugation of lipid peroxidation end products. Eur. J. Biochem. 268,2912–292312 Yan, L. J. <strong>and</strong> Sohal, R. S. (1998) Mitochondrial adenine nucleotide translocase ismodified oxidatively during aging. Proc. Natl. Acad. Sci. U.S.A. 95, 12896–1290113 Alin, P., Danielson, U. H. <strong>and</strong> Mannervik, B. (1985) 4-Hydroxyalk-2-enals aresubstrates for glutathione transferase. FEBS Lett. 179, 267–27014 Awasthi, Y. C., Zimniak, P., Awasthi, S., Singhal, S. S., Srivastava, S. K., Piper, J. T.,Chaubey, M., Petersen, D. R., He, N.-G., Sharma, R. et al. (1996) A new group ofglutathione S-transferases with protective role against lipid peroxidation. InGlutathione S-transferases: Structure, Function <strong>and</strong> Clinical Implications (Vermeulen,N. P. E., Mulder, G. J., Nieuwenhuyse, H., Peters, W. H. M. <strong>and</strong> van Bladeren, P. J.,eds.), pp. 111–124, Taylor & Francis, London15 Hubatsch, I., Ridderstrom, M. <strong>and</strong> Mannervik, B. (1998) Human glutathionetransferase A4-4: an Alpha <strong>class</strong> enzyme with high catalytic efficiency in theconjugation of 4-hydroxynonenal <strong>and</strong> other genotoxic products of lipid peroxidation.Biochem. J. 330, 175–17916 Clayton, J. D., Cripps, R. M., Sparrow, J. C. <strong>and</strong> Bullard, B. (1998) Interaction oftroponin-H <strong>and</strong> glutathione S-transferase-2 in the indirect flight muscles of Drosophilamelanogaster. J. Muscle Res. Cell Motil. 19, 117–12717 Chelvanayagam, G., Parker, M. W. <strong>and</strong> Board, P. G. (2001) Fly fishing for <strong>GST</strong>s: aunified nomenclature for mammalian <strong>and</strong> insect glutathione transferases. Chem. Biol.Interact. 133, 256–26018 Toung, Y.-P. S., Hsieh, T. <strong>and</strong> Tu, C.-P. D. (1993) The glutathione S-transferase Dgenes. A divergently organized, intronless gene family in Drosophila melanogaster.J. Biol. Chem. 268, 9737–974619 Beall, C., Fyrberg, C., Song, S. <strong>and</strong> Fyrberg, E. (1992) Isolation of a Drosophila geneencoding glutathione S-transferase. Biochem. Genet. 30, 515–52720 Singh, M., Silva, E., Schulze, S., Sinclair, D. A. R., Fitzpatrick, K. A. <strong>and</strong> Honda,B. M. (2000) Cloning <strong>and</strong> characterization of a new theta-<strong>class</strong> glutathione S-transferase (<strong>GST</strong>) gene, gst-3, from Drosophila melanogaster. Gene 247, 167–17321 Zimniak, P., Singhal, S. S., Srivastava, S. K., Awasthi, S., Sharma, R., Hayden, J. B.<strong>and</strong> Awasthi, Y. C. (1994) Estimation of genomic complexity, heterologous expression,<strong>and</strong> enzymatic characterization of mouse glutathione S-transferase m<strong>GST</strong>A4-4 (<strong>GST</strong>5.7). J. Biol. Chem. 269, 992–100022 Tang, A. H. <strong>and</strong> Tu, C.-P. D. (1994) Biochemical characterization of Drosophilaglutathione S-transferases D1 <strong>and</strong> D21. J. Biol. Chem. 269, 27876–2788423 Lee, H. C. <strong>and</strong> Tu, C.-P. D. (1995) Drosophila glutathione S-transferase D27:functional analysis of two consecutive tyrosines near the N-terminus. Biochem.Biophys. Res. Commun. 209, 327–33424 Toung, Y. P., Hsieh, T. S. <strong>and</strong> Tu, C.-P. D. (1991) The Drosophila glutathioneS-transferase 1-1 is encoded by an intronless gene at 87B. Biochem. Biophys.Res. Commun. 178, 1205–121125 Lougarre, A., Bride, J. M. <strong>and</strong> Fournier, D. (1999) Is the insect glutathioneS-transferase I gene family intronless? Insect Mol. Biol. 8, 141–14326 Feng, Y., Gunter, L. E., Organ, E. L. <strong>and</strong> Cavener, D. R. (1991) Translation initiation inDrosophila melanogaster is reduced by mutations upstream of the AUG initiatorcodon. Mol. Cell. Biol. 11, 2149–215327 Xiao, B., Singh, S. P., N<strong>and</strong>uri, B., Awasthi, Y. C., Zimniak, P. <strong>and</strong> Ji, X. (1999)Crystal structure of a murine glutathione S-transferase in complex with a glutathioneconjugate of 4-hydroxynon-2-enal in one subunit <strong>and</strong> glutathione in the other:evidence of signaling across the dimer interface. Biochemistry 38, 11887–1189428 Toung, Y. P. <strong>and</strong> Tu, C.-P. D. (1992) Drosophila glutathione S-transferases havesequence homology to the stringent starvation protein of Escherichia coli. Biochem.Biophys. Res. Commun. 182, 355–36029 Tang, A. H. <strong>and</strong> Tu, C.-P. D. (1995) Pentobarbital-induced changes in Drosophilaglutathione S-transferase D21 mRNA stability. J. Biol. Chem. 270, 13819–1382530 Rubin, G. M., Y<strong>and</strong>ell, M. D., Wortman, J. R., Gabor Miklos, G. L., Nelson, C. R.,Hariharan, I. K., Fortini, M. E., Li, P. W., Apweiler, R., Fleischmann, W. et al. (2000)Comparative genomics of the eukaryotes. Science 287, 2204–221531 Cavener, D. R. (1987) Comparison of the consensus sequence flanking translationalstart sites in Drosophila <strong>and</strong> vertebrates. Nucleic Acids Res. 15, 1353–136132 Kozak, M. (1999) Initiation of translation in prokaryotes <strong>and</strong> eukaryotes. Gene 234,187–208 2003 Biochemical Society


Cloning <strong>and</strong> expression of Drosophila glutathione S-transferases66933 Cheng, J. Z., Yang, Y., Singh, S. P., Singhal, S. S., Awasthi, S., Pan, S. S.,Singh, S. V., Zimniak, P. <strong>and</strong> Awasthi, Y. C. (2001) Two distinct 4-hydroxynonenalmetabolizing glutathione S-transferase isozymes are differentially expressed in humantissues. Biochem. Biophys. Res. Commun. 282, 1268–127434 Engle, M. R., Singh, S. P., N<strong>and</strong>uri, B., Ji, X. <strong>and</strong> Zimniak, P. (2001) Invertebrateglutathione transferases conjugating 4-hydroxynonenal: Ce<strong>GST</strong> 5.4 fromCaenorhabditis elegans. Chem. Biol. Interact. 133, 244–24835 Barrera, G., Pizzimenti, S., Muraca, R., Barbiero, G., Bonelli, G., Baccino, F. M.,Fazio, V. M. <strong>and</strong> Dianzani, M. U. (1996) Effect of 4-hydroxynonenal on cell cycleprogression <strong>and</strong> expression of differentiation-associated antigens in HL-60 cells. FreeRadicals Biol. Med. 20, 455–46236 Liu, W., Akh<strong>and</strong>, A. A., Kato, M., Yokoyama, I., Miyata, T., Kurokawa, K., Uchida, K.<strong>and</strong> Nakashima, I. (1999) 4-Hydroxynonenal triggers an epidermal growth factorreceptor-linked signal pathway for growth inhibition. J. Cell Sci. 112, 2409–241737 Pizzimenti, S., Barrera, G., Dianzani, M. U. <strong>and</strong> Brusselbach, S. (1999) Inhibition ofD1, D2, <strong>and</strong> A-cyclin expression in HL-60 cells by the lipid peroxidation product4-hydroxynonenal. Free Radicals Biol. Med. 26, 1578–158638 Cheng, J.-Z., Singhal, S. S., Saini, M., Singhal, J., Piper, J. T., Van Kuijk, F. J. G. M.,Zimniak, P., Awasthi, Y. C. <strong>and</strong> Awasthi, S. (1999) Effects of m<strong>GST</strong> A4 transfectionon 4-hydroxynonenal-mediated apoptosis <strong>and</strong> differentiation of K562 humanerythroleukemia cells. Arch. Biochem. Biophys. 372, 29–3639 Rinaldi, M., Barrera, G., Spinsanti, P., Pizzimenti, S., Ciafre, S. A., Parella, P., Farace,M. G., Signori, E., Dianzani, M. U. <strong>and</strong> Fazio, V. M. (2001) Growth inhibition <strong>and</strong>differentiation induction in murine erythroleukemia cells by 4-hydroxynonenal. FreeRadical Res. 34, 629–63740 Ranson, H., Rossiter, L., Ortelli, F., Jensen, B., Wang, X. L., Roth, C. W.,Collins, F. H. <strong>and</strong> Hemingway, J. (2001) Identification of a novel <strong>class</strong> ofinsect glutathione S-transferases involved in resistance to DDT in the malariavector Anopheles gambiae. Biochem. J. 359, 295–30441 Cheng, J.-Z., Sharma, R., Yang, Y., Singhal, S. S., Sharma, A., Saini, M. K.,Singh, S. V., Zimniak, P., Awasthi, S. <strong>and</strong> Awasthi, Y. C. (2001) Acceleratedmetabolism <strong>and</strong> exclusion of 4-hydroxynonenal through induction of RLIP76<strong>and</strong> h<strong>GST</strong>5.8 is an early adaptive response of cells to heat <strong>and</strong> oxidative stress.J. Biol. Chem. 276, 41213–4122342 Yang, Y., Cheng, J. Z., Singhal, S. S., Saini, M., P<strong>and</strong>ya, U., Awasthi, S. <strong>and</strong>Awasthi, Y. C. (2001) Role of glutathione S-transferases in protection againstlipid peroxidation. Overexpression of h<strong>GST</strong>A2-2 in K562 cells protects againsthydrogen peroxide-induced apoptosis <strong>and</strong> inhibits JNK <strong>and</strong> caspase 3activation. J. Biol. Chem. 276, 19220–1923043 Sohal, R. S. <strong>and</strong> Orr, W. C. (1998) Oxidative stress may be a causal factor insenescence. Age 21, 81–8244 Parkes, T. L., Elia, A. J., Dickinson, D., Hilliker, A. J., Phillips, J. P. <strong>and</strong>Boulianne, G. L. (1998) Ex<strong>ten</strong>sion of Drosophila lifespan by overexpressionof human SOD1 in motorneurons. Nat. Genet. 19, 171–17445 Seong, K.-H., Ogashiwa, T., Matsuo, T., Fuyama, Y. <strong>and</strong> Aigaki, T. (2001) Applicationof the gene search system to screen for longevity genes in Drosophila.Biogerontology 2, 209–21746 Kang, H. L., Benzer, S. <strong>and</strong> Min, K. T. (2002) Life ex<strong>ten</strong>sion in Drosophila by feedinga drug. Proc. Natl. Acad. Sci. U.S.A. 99, 838–84347 N<strong>and</strong>uri, B., Hayden, J. B., Awasthi, Y. C. <strong>and</strong> Zimniak, P. (1996) Amino acid residue104 in an alpha-<strong>class</strong> glutathione S-transferase is essential for the high selectivity<strong>and</strong> specificity of the enzyme for 4-hydroxynonenal. Arch. Biochem. Biophys. 335,305–31048 Armitage, P. <strong>and</strong> Berry, G. (1987) Statistical Methods in Medical Research. BlackwellScientific Publications, Oxford49 Felsenstein, J. (1993) PHYLIP (Phylogeny Inference Package), version 3.5c,Department of Genetics, University of Washington, Seattle, WAReceived 14 August 2002/8 November 2002; accepted 20 November 2002Published as BJ Immediate Publication 20 November 2002, DOI 10.1042/BJ20021287 2003 Biochemical Society

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!